Next Article in Journal
Aptamers as Smart Ligands for Targeted Drug Delivery in Cancer Therapy
Previous Article in Journal
Thermodynamic Correlation between Liquid–Liquid Phase Separation and Crystalline Solubility of Drug-Like Molecules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analysis of Time-Dependent Pharmacokinetics Using In Vitro–In Vivo Extrapolation and Physiologically Based Pharmacokinetic Modeling

1
Department of Life and Nanopharmaceutical Sciences, Kyung Hee University, 26, Kyungheedae-ro, Dongdaemungu, Seoul 02453, Republic of Korea
2
Division of Biopharmaceutics, College of Pharmacy, Kyung Hee University, Seoul 02447, Republic of Korea
3
Department of Integrated Drug Development and Natural Products, Kyung Hee University, Seoul 02447, Republic of Korea
*
Author to whom correspondence should be addressed.
Pharmaceutics 2022, 14(12), 2562; https://doi.org/10.3390/pharmaceutics14122562
Submission received: 2 November 2022 / Revised: 17 November 2022 / Accepted: 17 November 2022 / Published: 22 November 2022
(This article belongs to the Section Biopharmaceutics)

Abstract

:
SCR430, a sorafenib derivative, is an investigational drug exhibiting anti-tumor action. This study aimed to have a mechanistic understanding of SCR430’s time-dependent pharmacokinetics (TDPK) through an ex vivo study combined with an in vitro–in vivo extrapolation (IVIVE) and physiologically based pharmacokinetic (PBPK) modeling. A non-compartmental pharmacokinetic analysis was performed after intravenous SCR430 administration in female Sprague-Dawley rats for a control group (no treatment), a vehicle group (vehicle only, 14 days, PO), and a repeated-dosing group (SCR430, 30 mg/kg/day, 14 days, PO). In addition, hepatic uptake and metabolism modulation were investigated using isolated hepatocytes from each group of rats. The minimal PBPK model based on IVIVE was constructed to explain SCR430’s TDPK. Repeated SCR430 administration decreased the systemic exposure by 4.4-fold, which was explained by increased hepatic clearance (4.7-fold). The ex vivo study using isolated hepatocytes from each group suggested that the increased hepatic uptake (9.4-fold), not the metabolic activity, contributes to the increased hepatic clearance. The minimal PBPK modeling based on an ex vivo study could explain the decreased plasma levels after the repeated doses. The current study demonstrates the TDPK after repeated dosing by hepatic uptake induction, not hepatic metabolism, as well as the effectiveness of an ex vivo approach combined with IVIVE and PBPK modeling to investigate the TDPK.

Graphical Abstract

1. Introduction

Time dependency in pharmacokinetics (PK) refers to time-dependent pharmacokinetic changes by multiple dosing. On occasion, the term “time-dependent” may also be used for chronopharmacokinetics that demonstrate pharmacokinetic changes by the actual time of administration. In this study, time-dependent pharmacokinetics (TDPK) was defined as the dosing period time-dependent pharmacokinetics [1].
A structural sorafenib analog, SCR430, is a new drug candidate that exhibits anti-tumor activity by inhibiting p-STAT3 and inducing SHP-1 [2,3] (ClinicalTrials.gov Identifier: NCT04733521). SCR430’s TDPK potential was suggested in its toxicokinetic study. The AUC and Cmax decreased by 66% and 87%, respectively, in rats after repeated oral administration (30 mg/kg/day for 2 weeks).
The TDPK mechanism may vary. Metabolic enzyme auto-induction by repeated doses is a well-known mechanism, as reported in the cases of dexamethasone and rifampicin [4,5,6]. However, several biological factors besides metabolic enzymes govern drug elimination. For instance, the hepatic drug uptake and efflux transporters’ significance in the hepatic elimination process has been well documented, including safety, efficacy, and drug interactions [7,8,9]. Therefore, the auto-induction of such processes can also be the TDPK mechanism. Since hepatic elimination was expected to be the major elimination pathway of SCR430 by a preliminary PK study, this current study’s objectives were to investigate the underlying SCR430 TDPK mechanism by in vivo PK and hepatocyte ex vivo studies using in vitro–in vivo extrapolation (IVIVE) and physiologically based pharmacokinetic (PBPK) modeling [10,11].

2. Materials and Methods

2.1. Chemicals and Reagents

SCR430 (Figure 1) was kindly supplied by RaND Biosciences Inc. (Seongnam, Republic of Korea). Sorafenib, silicone oil, mineral oil, propylene glycol, kolliphor® HS 15, Williams’ medium E, and Hank’s balanced salts without calcium chloride and magnesium sulfate were purchased from Sigma-Aldrich, Inc. (St. Louis, MO, USA). Collagenase II was purchased from the Worthington Biochemical Corporation (Lakewood, NJ, USA). HPLC or MS-grade solvents used in the analysis were purchased from Sigma-Aldrich, Inc. (St. Louis, MO, USA).

2.2. Rat In Vivo PK Studies

Female Sprague-Dawley rats (8 weeks old, weighing between 190 and 205 g, Dahan Bio Link, Eumseong, Republic of Korea) were used for the in vivo PK studies. The female rats were used to evaluate urinary clearance by catheter insertion. The in vivo PK studies were conducted according to the schematic representation shown in Figure 1. Briefly, the rats were divided into three groups. The first group (control group) did not receive any treatment. The second group (vehicle group) and the third group (repeated-dosing group) were administered orally with a vehicle (50% Solutol HS15/50% polypropylene glycol 400 (1 mL/kg)) and 30 mg/kg of SCR430 dissolved in the vehicle once a day for two weeks, respectively. After the washout period of two days, the SCR430 that was dissolved in the vehicle was administered intravenously or intraperitoneally at a dose of 3 mg/kg for each group (n = 6/group). As time-dependent clearance (CL) was suggested in a previous study, the PK profile after the intravenous or intraperitoneal administration was investigated. Blood samples were collected in heparinized tubes based on a pre-designed schedule for up to 10 h. The blood samples were immediately separated into plasma and stored at −80 °C until analysis. Urine and bile samples were also collected to evaluate urinary and biliary excretions. The urine samples were collected for up to 10 h using urethral catheterization in each group (n =3). The bile samples were collected according to the designed schedule for up to 10 h using bile-duct cannulation from each group (n =3). The urine and bile samples were stored at −80 °C until analysis. All of the experimental procedures were reviewed and approved by the Committee on the Care and Use of Laboratory Animals, Kyung Hee University (KHSASP-19-440).

2.3. Non-Compartmental Analysis and Related PK Parameters

A pharmacokinetic analysis was performed by non-compartmental analysis using the PK-Solver® program [12]. The tissue-to-plasma partition coefficients (Kp) of the liver and kidney were measured 10 h after the IV administration using Equation (1), considering the remaining blood within the tissues [13]. The blood-to-plasma ratio (Rb) was determined in triplicate [14]. Briefly, the drug stock was spiked with fresh whole blood and plasma obtained from the control and repeated-dosing rats at a final concentration of 1 μM, and the samples were incubated at 37 °C for 1 h. Then, the blood samples were separated into plasma by centrifugation at 3000× g for 10 min at 4 °C. SCR430 and its metabolites in the plasma, bile, urine, and liver were analyzed using LC-MS/MS and UPLC-qToF-MS. The detailed procedures are described in Appendix A.

2.4. Determination of Fraction Unbound in Rat Plasma (fu,p), Hepatocyte Suspension (fu,inc,hepa), and Intrahepatocyte (fu,hepa)

Equilibrium dialysis (RED) assays were used to determine the free fraction of SCR430 in the rat plasma of the control and repeated-dosing groups and the hepatocyte suspension of the control group. An RED device (Thermo Fisher Scientific, Rockford, USA) was prepared according to the manufacturer’s guidelines. A total of 300 μL of plasma and dead hepatocyte suspension containing 5 μM SCR430 and 0.1% DMSO and 500 μL of phosphate-buffered saline (PBS) with solutol® (0.01% v/v in PBS) were loaded into the sample (donor) chamber and buffer (receiver) chamber, respectively [15]. The samples were incubated at 37 °C on an orbital shaker at 300 rpm for 16 h. At the end of the incubation, the samples from both chambers were collected. Then, an equal volume of each matrix was added for matrix matching to improve the analytical efficiency. Finally, a liquid–liquid extraction (LLE) was performed on the samples using methyl tert-butyl ether and quantified by LC-MS/MS analysis. The fu,inc,hepa of the repeated-dosing group was assumed to be equal to that of the control group because of the non-specific binding of SCR430 in the hepatocyte suspension.
The fu,hepa was calculated from the reciprocal of the hepatocyte to the medium concentration ratio of the suspended hepatocyte isolated from the control and repeated-dosing groups at 4 °C by an oil-spin method to be described later under the assumption that the cellular binding is not temperature-dependent and that the active transport is entirely abolished on ice [16].
f u , h e p a = 1 A c e l l V c e l l C b u f f e r
where Vcell is the rat hepatocyte volume (3.68 ± 1.37 μL/106 cells) [17], Acell represents the amount of SCR430 in the hepatocytes, and Cbuffer is the SCR430 concentration in the buffer layer. The incubation proceeded for 120 min at 4 °C.

2.5. Assessment of Intrinsic Metabolic Clearance (CLint,met), Hepatic Uptake Clearance (PSinf), and Liver-to-Plasma Unbound Drug Concentration Ratio at Steady State (Kpuu,ss) Using Isolated Hepatocytes

Fresh rat hepatocytes were obtained from the control and repeated-dosing groups using the two-step collagenase liver perfusion method, described by Seglen et al. [18].
The intrinsic metabolic clearance (CLint,met) was extrapolated from the in vitro intrinsic metabolic clearance (CLint,met,vitro), and the CLint,met,vitro was determined by the conventional hepatocyte metabolic stability assay, which measures the parent drug depletion in the whole incubation medium [19]. Briefly, a drug stock (1 mM) dissolved in dimethyl sulfoxide and freshly isolated hepatocytes (1 × 106 cells/mL) obtained from rats in the control, vehicle, and repeated-dosing groups were pre-incubated at 37 °C for 15 min. The metabolic reaction was initiated by adding 1 μL of the pre-incubated drug into 999 μL of the pre-incubated hepatocyte suspension in a buffer medium to achieve a final concentration of 1 μM containing 0.1% (v/v) dimethyl sulfoxide. At various time intervals (0, 5, 10, 20, 30, and 60 min), aliquots (50 μL) were transferred to 1 mL of ice-cold methyl tert-butyl ether solution. The extraction process and LC-MS/MS analysis were performed as described in Appendix A. Based on the predicted degradation half-life from the peak area ratio of the time profile data, the CLint,met,vitro (microliters per minute per 106 cells) was calculated.
The CLint,met was determined using the following equation from CLint,met,vitro:
C L i n t ,   m e t = C L i n t ,   m e t ,   v i t r o f u ,   i n c ,   h e p a   × p S F
The physiologic scaling factor for the rats (pSF) was 108 million cells/g liver and 36 g liver/kg body weight [20].
The PSinf, the sum of PSinf,act + PSdiff, was extrapolated from the in vitro hepatocyte uptake clearance (PSinf,hep), and the PSinf,hep was determined by an integration plot analysis by an oil-spin uptake assay using the suspended hepatocytes isolated for the control and repeated-dosing groups using the following Equations (3) and (4) [21,22].
The uptake assay was performed at 37 °C (for the active and passive uptakes) and at 4 °C (on ice for the passive uptake) to discriminate between the active and passive processes. The hepatic uptake was initiated by adding buffers containing the drug (2 μM in 0.2% DMSO) to the suspended hepatocytes in the buffer at an equal volume, resulting in the final concentration of the drug (1 μM in 0.1% DMSO) and a hepatocyte concentration of 0.5 × 106 cells/mL. After 10, 45, 60, 90, and 120 s, aliquots (100 μL) were removed from each tube and transferred to a silicon layer tube that consisted of two layers; the upper layer included 100 µL of oil mixture (silicon oil (81.7)/mineral oil (18.3), v/v, with a density of 1.015 g/mL) and the bottom layer included 50 µL of 1 M sucrose. Then, centrifugation was performed at 7000× g for 10 s to distinguish between the hepatocytes and the medium buffer. After the centrifugation, the upper- and lower-layer samples were separated and kept in a −80 °C deep freezer until further analysis.
The PSinf,hep was calculated from the slope of the integration plot and used for the extrapolation.
X h e p   ( t ) C b u f f ( t ) = P S   i n f , h e p × A U C ( 0 t ) ,   b u f f C b u f f ( t ) + V 0
P S i n f = P S   i n f , h e p f u ,   b u f f   × p S F
where Xhep(t), Cbuff(t), AUC(0−t),buff, fu,buff, and V0 represent the amount of the drug that was taken into the hepatocytes, the concentration of the drug in the buffer layer after centrifugation, the area under the drug concentration–time curve in the buffer layer, the unbound fraction of the drug in the buffer layer, which was assumed to be 1, and the initial volume of distribution at the initial stage, respectively. The Xhep(t)/Cbuff(t) value was plotted against the AUC(0−t),buff/Cbuff(t) value, and the PSinf,hep (µL/min/106 cells) was estimated from the initial slope obtained from the linear plot.
The Kpuu,ss was determined using rat hepatocytes isolated from the control and repeated-dosing groups. A temperature-dependent steady-state uptake method by oil-spin uptake assay with 4% rat serum at 37 °C and 4 °C was used [16,22].
The Kpuu,ss was calculated using the following Equation (5):
K p u u ,   s s   = C u ,   h e p , s s   ( 37   ° C ) C u ,   m e d i a , s s   ( 37   ° C ) = C h e p , s s   ( 37   ° C ) × f u , h e p a C m e d i a , s s   ( 37   ° C ) × f u , m e d i a = K p ( 37   ° C ) K p ( 4   ° C )
The Kp was determined by the ratio of the hepatocyte concentration and the medium concentration at 20 min (37 °C) and 120 min (4 °C).

2.6. Calculation of In Vivo Overall Intrinsic Clearance (CLint,all) and Overall Hepatic Clearance (CLh)

An IVIVE based on extended clearance concepts was applied to predict the CLint,all [11].
The CLint,all was defined as a function of the intrinsic metabolic CL (CLint,met), the hepatic diffusional passive (PSdiff) and hepatic active uptakes (PSinf,act), and the efflux (PSeff,act) permeability. Biliary efflux was not counted because SCR430 was detected at a trace level in the bile.
C L i n t ,   a l l   = C L i n t ,   m e t   × ( P S i n f ,   a c t + P S d i f f ) ( P S e f f ,   a c t + P S d i f f + C L i n t ,   m e t   )
The CLint,all was calculated in three ways [16,22,23].
Method 1: Determination of the CLint,all by assuming the intrinsic metabolic clearance dependent (CLint,met).
Method 1 assumed that the active transport was negligible (PSinf,act and PSeff,act « PSdiff) and that the hepatic diffusional passive permeability greatly exceeded the intrinsic metabolic clearance (PSdiff » CLint,met) in Equation (6) [22].
C L i n t ,   a l l   C L i n t ,   m e t  
Method 2: Determination of the CLint,all by assuming the hepatic uptake clearance (PSinf).
If the CLint,met is much greater than the overall total efflux clearance (PSeff,act + PSdiff) in Equation (6), the CLint,all is dominated by the overall hepatic uptake clearance (PSinf = PSinf,act + PSdiff) as the rate-limiting step [16,24].
C L i n t ,   a l l   P S i n f
Method 3: Determination of the CLint,all based on the liver-to-plasma unbound drug concentration ratio at a steady state (Kpuu,ss).
The right part of Equation (6) for the CLint,all, (PSinf,act + PSdiff)/(PSeff,act + PSdiff + CLint,met) corresponds to the Kpuu,ss, regardless of the assuming rate-limiting step [22,25,26].
C L i n t ,   a l l   = C L i n t ,   m e t × K p u u , s s

2.7. Determination of Fraction Unbound in Rat Plasma (fu,p), Hepatocyte Suspension (fu,inc,hepa), and Intrahepatocyte (fu,hepa)

The hepatic clearance (CLh) was predicted from the following equations based on a dispersion model using the CLint,all derived from the above three methods (Method 1, 2, and 3) [24]:
C L h = Q h   × ( 1 F h ) × R b
F h   = 4 a ( 1 + a ) 2 × exp ( a 1 2 D N ) ( 1 a ) 2 × exp ( a + 1 2 D N )
a = ( 1 + 4 R N × D N ) 1 2
R N = f u , b l o o d × C L i n t ,   a l l   Q h  
The DN, dispersion number, was set to 0.17 [27], and Qh was 3.69 L/h/kg [28].

2.8. Accuracy, Precision, and Bias Assessments of the Model Predictions

The accuracy of the estimated CLh was calculated by the fold error method, based on whether the ratio of the model-predicted values fell within 3-fold of the observed in vivo CLh values in the rats [16].
f o l d   e r r o r = 10 log 10 p r e d i c t e d   C L h o b s e r v e d   C L h  
The following equation uses the absolute average fold error (AAFE) to measure the bias:
A A F E = 10 1 n   log ( p r e d i c t e d   C L h o b s e r v e d   C L h   )

2.9. The Minimal PBPK (mPBPK) Model Construction for the Control and Repeated-Dosing Groups

The liver was considered the major elimination pathway, considering low biliary and renal clearance data. Thus, the mPBPK containing the liver and a single-adjustment compartment was constructed to predict the plasma and liver concentrations of the control and repeated-dosing groups (Figure 7). To emulate the dispersion model, the liver compartment was divided into five units, which were composed of extracellular and hepatocellular units linked by hepatic blood flow [29]. The mass-balanced differential equations for the mPBPK model referring to the liver are shown in Appendix A. The kinetic input parameters for the mPBPK modeling of the control and repeated-dosing groups are summarized in Table 3. The fu,hepa values of the control and repeated-dosing groups were assumed to be equal to the fu,hepa of the control and repeated-dosing groups. The fu,b was calculated from the fu,p, and the Rb.
The mPBPK model was coded using Berkeley Madonna (version 8.3.23, University of California at Berkeley, CA, USA). The performance of the constructed PBPK model was assessed using the determination coefficient (R2) and the mean absolute percentage error.

2.10. Sensitivity Analysis

Using Berkeley Madonna, a local sensitivity analysis was performed to determine the parameters that have the most crucial influence on the plasma or liver AUC for SCR430. The differences in the AUC were estimated after each parameter was increased by 10%. Equation (16) was used to compute the normalized sensitivity coefficient (NSC) [30,31,32,33], where r and r′ represent the initial value and the modified value of the plasma or liver AUC by the 10% increase in the interest parameter (e.g., fu,hepa, PSinf,act), respectively. Additionally, p and p′ represent the initial value and the 10% increased value of the interest parameter, respectively. The relative impact of each parameter on the plasma and liver AUC was categorized as follows: low: |NSC| < 0.2; medium: 0.2 ≤ |NSC| < 0.5; high: 0.5 ≤ |NSC| [34].
NSC   = r r r × p p p

2.11. Comparison of the Model Predictions Using Monte Carlo Simulation and Observed Data of Plasma and Liver Concentrations

The Monte Carlo simulation (n = 1000) was employed to assess the impact of uncertainties in the parameter values, including the hepatic uptake or metabolic process and the unbound fraction in the liver (fu,hepa) on the plasma and liver concentrations in the control and repeated-dosing groups. Lognormal distributions were assumed for the model parameters, including fu,hepa, PSinf,act, PSdiff, and CLint,met.
The probabilistic distributions of the lognormally distributed parameter values were defined by their mean values and the variability of the in vitro data derived from the uptake, metabolic stability, and binding assays using the rat hepatocytes. First, the skewed parameters with a lognormal distribution were transformed into a normally distributed variable ω, with the mean (μx) and the standard deviation (σx) using the equation shown as follows [35]:
μ ω = ln ( μ x   2   /   σ x   2 + μ x   2
σ ω = ln ( 1 + σ x   2 μ x   2 )
Lognormal   ( μ x ,   σ x ) = exp ( Normal   ( μ ω ,   σ ω )
where μx and σx are the mean and the standard deviation for the lognormal distribution of the parameters derived from the in vitro data, respectively; moreover, μω and σω represent the mean and the standard deviation following the conversion to a normal distribution, respectively [36].
The Monte Carlo simulations were implemented within Berkeley Madonna as described by Li et al. [35]. Intravenous doses of 3 mg/kg of the SCR430 were simulated for the plasma and liver concentrations of the control and repeated-dosing groups, respectively. Excel and GraphPad Prism 6.0 (GraphPad Software Inc., San Diego, CA, USA) were used to calculate the 5th, 20th, 35th, 50th, 65th, 80th, and 95th percentiles of the liver and plasma concentrations in the control and repeated-dosing groups.

2.12. Gene Expression Analysis Using Quantitative PCR (qPCR)

The Rat pregnane X receptor (rPXR), the rat farnesoid X receptor (rFXR), the rat Cytochrome P450 3a2 (rCyp3a2), the rat organic cation transporter1 (rOct1), and the rat organic anion transporting polypeptides (rOatp1a1, rOatp1a4, and rOatp1b2) were chosen as the representative genes related to hepatic auto-induction or elimination. The mRNA expression of the genes was monitored using real-time qPCR in the liver that was extirpated from the rats (n = 3) and an in vivo PK study was completed in the control and repeated-dosing groups. According to the manufacturer’s instructions, the total RNA was extracted from the liver using RNAiso (Takara, Tokyo, Japan) and used for the qPCR [37]. The primer sequences and PCR conditions for the qPCR are summarized in Appendix B, Table A1. The result of integrating the standard deviation of the ΔΔCT value was used to calculate the fold difference [38].

3. Results

3.1. TDPK of SCR430

The SCR430’s mean plasma concentration–time curves after the intravenous bolus administration dosing at 3 mg/kg in the control, vehicle, and repeated-dosing groups indicated SCR430’s TDPK nature (Figure 2a). The repeated-dosing group showed lower plasma concentrations with a more rapid elimination phase than the control and vehicle groups. In contrast, the control and vehicle groups’ plasma profiles were similar. The PK parameters that were estimated by the non-compartmental analysis indicated the SCR430’s time-dependent CL (Figure 2c–e, and Table 1). The repeated-dosing group’s AUC0–∞ and CL showed a statistically significant difference in the AUC0–∞ and CL compared to the control and vehicle groups, whereas the volume of distribution at the steady state (Vdss) was similar. An approximately 4-fold decrease in the AUC0–∞ (p = 0. 0013) and an increase in the CL (p < 0.001)) were observed. In contrast, the vehicle group’s AUC0–∞, CL, and Vdss were not statistically different from those of the control group (p > 0.05). Additionally, similar results were observed after an IP administration, where TDPK was more clearly identified after the IP administration (Figure 2b). The plasma concentration of the repeated-dosing group was much lower than that of the control and vehicle groups, and an 8-fold decrease in the AUCIP,0–10 was observed in the repeated-dosing group compared to the control group (p < 0.001). The biliary and renal clearances were negligible and less than 0.1% compared to the total clearance in all of the groups (Figure 3a, Table 1). SCR430’s biliary clearance did not show any statistical differences between the groups (Figure 3b). However, the urinary clearance was higher in the repeated-dosing group (Figure 3c). The Kp of the liver in all of the groups presented no significant differences. The Kp of the kidney was higher in the repeated-dosing group than in the control and vehicle groups (Table 1). The Rb values were calculated as 3.95, 4.58, and 4.53 in the control, vehicle, and repeated-dosing groups, respectively. There were no significant differences in the groups (Table 1).

3.2. Metabolite Identification and Quantification of SCR430

SCR430’s dominant metabolites were predicted as hydroxyl metabolites by in silico metabolite profiling using GLORY [41]. From three candidate metabolites for the SCR430 site (Figure 4a), two structurally different hydroxylated metabolite types at m/z 446.052 [M-OH] were confirmed using UPLC-qToF-MS in the plasma, liver tissue, bile acid, and hepatocyte samples (Figure 4b). The identified metabolites were not practically analyzed in urine samples. Due to chemical standard unavailability for the identified metabolites, the equivalent hydroxylated metabolite concentration was predicted using LC-MS/MS on the assumption that the linearity of the metabolite MS response and the concentration is equivalent to the parent drug. The equivalent hydroxylated metabolite amounts excreted in the bile for up to 10 h from each group were converted into percent of dose. Approximately ten times more metabolites were generated in the repeated-dosing group than in the control and vehicle groups (Figure 4c).

3.3. Low Fraction Unbound of SCR430 in the Plasma and Liver

The fu,p, fu,inc,hepa, and fu,hepa values were low and determined to be 0.0078, 0.032, and 0.004 for the control group, respectively (Table 1). These low values of fraction unbound may be due to SCR430’s high lipophilicity (clogP = 6.15, Appendix B, Table A2). The fu,p and fu,inc,hepa values predicted by an in silico prediction based on physicochemical properties (Appendix B, Table A2) was 0.0066 [42] and 0.024 [39], similar to the observed values, which were 0.0078 and 0.032.
Interestingly, the repeated-dosing group’s fu,hepa value was 2.5-fold higher than the control group’s value, whereas the control group’s fu,p was similar to the vehicle and repeated-dosing groups’ values. Although the reason was yet unclear, this difference clearly explained the liver concentration difference in the repeated-dosing group, which will be described later.

3.4. Similar CLint,met between the Control, Vehicle, and Repeated-Dosing Groups

The difference in the CLint,met by the hepatic metabolism between the three groups was assessed by a conventional metabolic stability assay using the suspended rat hepatocytes.
A relatively slow depletion of SCR430 was observed with monophasic kinetics. The CLint,met of the control, vehicle, and repeated-dosing group was 1.32, 1.21, and 1.69 µL/min/106 cells, respectively (Figure 5a–c). No significant difference in the CLint,met was observed between the groups (Table 2).

3.5. Increased PSinf,hep and Kpuu,ss in the Repeated-Dosing Group

The time-dependent PSinf,hep and Kpuu,ss potentials were investigated in the suspended hepatocytes isolated from the control and repeated-dosing groups’ livers (Figure 5d–g). The PSinf,hep calculated from the initial uptake clearance by an integration plot was significantly increased in the repeated-dosing group compared to the control group (p = 0.0263, Table 2). The control and repeated-dosing groups’ Kpuu,ss values were determined by a temperature method at 4 °C and 37 °C and were 0.11 and 1.7, respectively (Table 2). An approximately 15-fold Kpuu,ss value increase in the repeated-dosing group was observed compared to the control group (p < 0.001).

3.6. Comparison of CLh Determined by Three IVIVE Methods and In Vivo Clearance in the Control and Repeated-Dosing Groups

The CLh was predicted by the following three IVIVE methods: (1) intrinsic metabolic clearance, (2) hepatic uptake clearance, and (3) liver-to-plasma unbound drug concentration ratio at steady state. They were determined in the control and repeated-dosing groups and used for the CLh prediction (Table 2). The observed total in vivo CL was compared with the predicted CLh by the three IVIVE methods (Figure 6), as SCR430’s renal clearance was negligible. The CLh obtained by Method 2 were clearly explained by the observed in vivo CL within a 3-fold (Table 2) and were centered along the unity line (Figure 6b). However, Method 1 failed to explain the difference in the in vivo CL between the control and repeated-dosing groups (Figure 6a). Furthermore, there was no statistical difference in the predicted CLh between the control and the repeated-dosing groups. Method 3 showed a predicted value that was slightly overestimated over 3-fold for the control group, and an increased CL in the repeated-dosing group was partially predicted (Figure 6c). Considering the uncertainty caused by the fu,inc,hepa (0.032) in calculating the CLint,met [43], the other values of fu,inc,hepa (0.024, value from the in silico prediction [39], 0.088, extrapolation from fu,p [40], and unity) were also used in Methods 1 and 3, but similar results were observed. Method 1 still did not reflect the change in the in vivo CL in the repeated-dosing group. In the case of Method 3, the absolute difference between the predicted and observed values was still larger than Method 2, although the change in the repeated-dosing group was partially reflected.
When the data were analyzed by the AAFE method, the 3-fold error range was between 0.37 and 2.8. The AAFE value for the CLh calculated by Method 1 in the control group was 2.4, indicating a disagreement between prediction and observation. The AAFE values for the CLh calculated by Methods 2 and 3 were 0.73 and 0.26, respectively, for the control group and 0.53 and 0.99, respectively, for the repeated-dosing group. These results suggested that the hepatic uptake was a more critical factor in explaining the SCR430 TDPK.

3.7. An mPBPK Model Could Predict the Differences in Plasma and Liver Concentrations in Control and Repeated-Dosing Groups

An mPBPK model has been developed to elucidate how changes in the hepatic uptake by auto-induction affect the overall plasma and liver concentrations (Figure 7). The volumes of the single-adjustment and systemic compartments and the rate constant to the single-adjustment compartment (kin) and from (kout) were fitted to the actual plasma concentration profile. The PSinf was subdivided into passive diffusion clearance (PSdiff) and active hepatic uptake clearance (PSinf,act) for the liver modeling. The PSinf,act and PSdiff were calculated using IVIVE Method 2, and the empirical scaling factor (1.6) was multiplied to match the in vivo total CL based on a sensitivity analysis, which will be described later. The PSinf,act and PSdiff were 2.5 L/h/kg and 0.8 L/h/kg, respectively, for the control group and 9.2 L/h/kg and 0.6 L/h/kg, respectively, for the repeated-dosing group (Table 3). The other mPBPK model parameters that were used are listed in Table 3.
The Monte Carlo simulation using the constructed mPBPK model showed a similar profile to the control and repeated-dosing groups’ observed plasma and liver concentrations. The observed concentration data in the plasma and liver for both of the groups were within the 95th percentile of the predicted data generated by the Monte Carlo simulations (Figure 8a–d). The prediction score with mean absolute percentage errors for the control and repeated-dosing groups showed an acceptable prediction (within 50%), and the determination coefficient (R2) was 0.85 for the control group and 0.93 for the repeated-dosing group. Especially, the time-dependent CL and liver concentration differences in the repeated-dosing group were reasonably predicted by the change in the PSinf,act and PSdiff in the mPBPK model. The decreased liver concentration in the repeated-dosing group could be explained by the increased fu,hepa in the model.

3.8. Sensitivity Analysis

From all of the parameters related to hepatic elimination, PSinf,act, PSdiff, CLint,met, and fu,hepa exhibited high absolute values of the normalized sensitivity coefficient (>0.2), and these parameters were used for a sensitivity analysis (Figure 9).
As expected from previous results, the plasma concentration was the most sensitive to the PSinf,act and less sensitive to the PSdiff, CLint,met, and fu,hepa. In contrast, the most influential factors for the liver concentration were CLint,met and fu,hepa.

3.9. The mRNA Expression Differences between the Three Groups

The mRNA expression change, which is relevant for drug elimination in rat livers and kidneys, is summarized in Figure 10. The repeated-dosing group resulted in a significant increase in the expression levels of rOct1 in the liver (p = 0.0082). Additionally, the values of rPXR and rFXR tended to decrease, with no statistically significant differences.

4. Discussion

The auto-induction potential is a critical issue in the drug development stages and should be discussed for its quantitative impact on the PK parameters of the drug itself or co-administered drugs [45]. Until now, most auto-inductions were reported to be related to CYPs enzyme induction [46,47]. Only a limited number of studies are reported on hepatic transporter auto-induction along with metabolic enzyme auto-induction [48,49]. In this study, the underlying SCR430 TDPK mechanism in rats was investigated in terms of hepatic uptakes, metabolic activity, and binding in the matrix using IVIVE and PBPK modeling for the control and repeated-dosing groups.
The differences in the in vivo PK properties of the three groups following IV and IP administration at a 3 mg/kg dose were compared. The vehicle group did not alter the drug’s pharmacokinetic properties compared to the control group. However, a considerable decrease in the plasma concentration was shown in the repeated-dosing group compared to the control group (Figure 2a,b). The results indicate a time-dependent CL in the repeated-dosing group. Interestingly, the TDPK was more significant after the IP administration, which was thought to be due to the difference in the first-pass effect according to the administration route.
In contrast, the Vdss showed no significant differences in the groups (Table 1). The SCR430 elimination into the bile and urine as a parent form was negligible. The majority of the SCR430 was excreted into the bile in a hydroxylated metabolite form. This finding suggested that SCR430 appears to be cleared primarily by the liver and that the drug-induced induction in the hepatic elimination may be the primary SCR430 TDPK mechanism.
However, a SCR430 with a high molecular weight (≥400 Da), less likely biliary excretion, and high lipophilicity was considered cleared by the active transporter-mediated hepatic uptake process according to the extended clearance classification system [11,50]. Therefore, an ex vivo approach using isolated hepatocytes from the control and repeated-dosing groups was used in this study based on the extended clearance concept to discriminate the contribution of metabolic enzymes and hepatic transporters. The extended clearance concept is known to be useful in investigating the complicated interactions between metabolism and transport [10,11].
Interestingly, a statistically significant difference in the CLint,met was not observed between the three groups (Figure 5a–c), and the CLh calculated using the IVIVE Method 1 (metabolic-dependent) could, therefore, not explain the time-dependent CL in the repeated-dosing group. However, the PSinf in the repeated-dosing group was larger than in the control group, and the CLh calculated using the IVIVE Method 2 (uptake-dependent) could explain the SCR430’s time-dependent CL. Method 3 was expected to predict changes in the in vivo CL, similar to Method 2. Method 3 could partially predict the induction of the CL in the repeated-dosing group, but the prediction precision was lower than in Method 2. Therefore, the experimental observations in this study suggested that SCR430 may induce uptake transporters in rats’ livers, which could be the primary TDPK mechanism. The increased hepatic uptake may also explain the increased Kp of the kidney in the repeated-dosing group because hepatic uptake transporters, such as organic cation transporters, are also expressed in the kidney.
The mPBPK models for SCR430 also suggested that the inducible SCR430 hepatic uptake is most likely a contributor to the decreased plasma systemic exposure observed after multiple SCR430 dosing. The SCR430 plasma concentration–time profiles simulated by the mPBPK model, which incorporate parameters reflecting uptake induction, were in good agreement with the observed data in both of the groups (Figure 8). The Monte Carlo simulation and the sensitivity analysis results also demonstrated that the variations in the PSinf,act may be the primary explanations for the decrease in the systemic exposure following the repeated SCR430 administration. Moreover, the increased fu,hepa in the repeated-dosing group determined from in vitro hepatocyte assays could predict in vivo liver concentrations within a 2-fold range.
The mRNA expression level changes may indicate an increased PSinf,act. The rOct 1 expression showed about a 2.4-fold increase in the liver, whereas there were no significant changes in the metabolic enzyme’s mRNA expression. Interestingly, the nuclear receptors rPXR and rFXR were reported to suppress OCT1 expression [51,52]. The rPXE and rFXR expression levels showed a decreased tendency (Figure 10), although statistical significance was not found. However, the meaning of the mRNA expression results should be interpreted carefully because SCR430’s substrate specificity for rat transporters is not available and a limited number of genes were tested.
Unfortunately, we do not have a clear explanation for the change in the fu,hepa, including PSdiff, in the repeated-dosing group. However, the mechanism might be related to the SCR430 target considering that passive permeation through lipid membranes depends on lipid composition, and fu,hepa is known to depend on non-specific binding to the cellular phospholipids [53,54]. SHP-1 is already known to play a crucial role in lipid metabolism in the liver [55]. Further studies on the mechanism for changes in the PSinf,act, PSdiff, and fu,hepa are likely needed.
As the chemical compound of interest has become more metabolically stable recently, transporter-mediated CL processes have become the primary factors behind clearance [25,56,57,58]. For instance, it has been demonstrated that the intrinsic hepatic uptake CL in the case of statins, which was determined in human hepatocyte suspension, was far more accurate for predicting the in vivo clearance than extrapolating directly from the intrinsic metabolic CL [16,59]. Therefore, our results, TDPK by auto-induction of the hepatic uptake process, and approach, mechanistic analysis using IVIVE and PBPK, are worth carefully considering in new drug development processes.

5. Conclusions

This study’s results revealed that SCR430 showed time-dependent CL after repeated doses and was related to the hepatic uptake process auto-induction, not to metabolic activity. This is the first report on TDPK induced by the transport process without metabolic involvement.
The ex vivo approach combined with IVIVE and PBPK, established in our study, could help mechanistically clarify complex auto-induction mechanisms in the drug development stage.

Author Contributions

Conceptualization, Y.-J.L. and M.-C.K.; methodology and investigation, M.-C.K.; writing, Y.-J.L. and M.-C.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the RaND Biosciences Inc. [KHU20192414]; the Basic Science Research Program through the National Research Foundation of Korea, funded by the Ministry of Education [2019R1A2C1011350].

Institutional Review Board Statement

The animal study protocol was approved by the Committee on the Care and Use of Laboratory Animals, Kyung Hee University (KHSASP-19-440).

Acknowledgments

We thank H.-K. Baek, J.-W. Lee, T.-H. Park, H.-J. Lee, H.-W. Lee and H.-S. Park for their technical assistance.

Conflicts of Interest

Y.-J.L. received research grants from the RaND Biosciences Inc. (KHU20192414). The funder had no role in the design of the study; in the collection, analyses, or interpretation of the data; in the writing of the manuscript; or in the decision to publish the results.

Appendix A

Appendix A.1. Analytical Method

All of the samples were processed with liquid–liquid extractions. Aliquots (50 μL) from each sample were transferred into 2 mL microfuge tubes, and the reaction was terminated by adding 1 mL of ice-cold methyl tert-butyl ether solution and 50 μL sorafenib, an internal standard. After vortex mixing, followed by centrifugation at 7000× g for 10 min, 900 μL of the supernatants was evaporated by drying under a stream of nitrogen gas. The residues were reconstituted with 100 μL of the mobile phase for the LC-MS/MS analysis. An Acquity UPLC (Waters, Milford, MA, USA) with a Waters Quattro micro™ API mass spectrometer (Waters, Milford, MA, USA) and a Meteoric Core C18 100A, 100 × 2.0 mm, 2.7 µm (YMC, Kyoto, Japan) were used for the study.
An isocratic mobile phase consisting of an 80:20 mixture of 0.01% formic acid in acetonitrile (ACN) and 0.01% formic acid in distilled water was delivered at a flow rate of 0.2 mL/min. Quantification was achieved using electrospray ionization-MS/MS detection in the positive ion mode for the SCR430, its oxidized metabolite, and sorafenib. The optimal instrumental parameters were as follows: capillary voltage of 3.5 kV, cone voltage of 40 V, source temperature of 100 °C, and desolvation temperature of 300 °C. Nitrogen was used as the desolvation gas and the cone gas, with flow rates of 500 and 50 L/h, respectively. The collision energy was 20 V. Detection of the ions was carried out in the multiple reaction-monitoring mode, and the settings are shown in Appendix B, Table A3. The identification of the SCR430 metabolites in the plasma, liver tissue, bile, and hepatocytes was achieved using UPLC-quadrupole time-of-flight MS (qToF-MS), SYNAPT G2-Si HDMS, (Waters, Milford, MA, USA). The settings of the qToF-MS were as follows: MS scan range, 60–1400 m/z (ESI (-) mode, capillary voltage of 2 kV, cone voltage of 10 V, source temperature of 100 °C, and desolvation temperature of 350 °C.

Appendix A.2. The Mass-Balanced Differential Equation for the Liver Compartments

A permeability-limited model was used for each compartment, incorporating the active transporter-mediated uptake (PSinf,act), the passive diffusion clearance (PSdiff), and the intrinsic metabolic clearance CLint,met. The corresponding differential equations referring to the liver were as follows:
Extracellular compartment 1:
V h e 5 × d C h e , 1 d t = Q h × ( C s y s C h e , 1 ) P S i n f ,   a c t + P S   d i f f 5 × f u ,   b × C h e , 1 + P S d i f f 5 × f u ,   h e p a × C h c , 1  
Extracellular compartment 2~5 (i = 2~5):
V h e 5 × d C h e , i d t = Q h × ( C h e , i 1 C h e , i ) P S i n f ,   a c t + P S   d i f f 5 × f u ,   b × C h e , i + P S d i f f 5 × f u ,   h e p a × C h c , i  
Hepatocellular compartment 1~5 (i = 1~5):
V h c 5 × d C h c , i d t = P S i n f ,   a c t + P S d i f f 5 × f u , b × C h e , i P S d i f f + C L i n t , m e t 5 × f u , h e p a × C h c , i
where Che, Chc, Csys, and Vhe are the concentration in the liver extracellular compartment, the concentration in the hepatocellular compartment, the concentration in the systemic compartment, and the volumes of the extracellular compartments, respectively (Table 3).

Appendix B

Table A1. Primer sequences and PCR conditions used for the qPCR.
Table A1. Primer sequences and PCR conditions used for the qPCR.
GeneForward Primer
(5′-3′)
Reverse Primer
(5′-3′)
Product Size (bp)PCR Conditions
rGapdhAACGACCCCTTCATTGACTCCACGACATACTCAGCAC191D: 95 °C for 10 s,
A: 57 °C for 30 s,
E: 72 °C for 30 s
rPXRTCCACTGCATGCTGAAGAAGAACCTGTGTGCAGGATAGGG187D: 95 °C for 10 s,
A: 56 °C for 30 s,
E: 72 °C for 40 s
rFXRCCAACCTGGGTTTCTACCCCACACAGCTCATCCCCTTT183D: 95 °C for 10 s,
A: 56 °C for 30 s,
E: 72 °C for 40 s
rCyp3a2CGACTTGGAACCCATAGACCATGTCAAATCTCCCTAAGCC117D: 95 °C for 10 s,
A: 58 °C for 30 s,
E: 72 °C for 30 s
rOatp1a1
(Oatp1)
ACCTGGAACAGCAGTATGGAAAAACCGATAGGCAAAATGCTAGGTAT163D: 95 °C-5 to 10 s,
A: 56 °C for 6 s,
E: 72 °C for 6 s
rOatp1a4
(Oatp2)
TGTGATGACCGTTGATAATTTTCCATTCTCCACATATAGTTGGTGCTGAA81D: 95 °C-5 to 10 s,
A: 56 °C for 10 s,
E: 72 °C for 6 s
rOct1CTCTGGCTACAGGAGAACGACCCTGGTACAGCACAGCACAA401D: 94 °C for 30 s,
A: 60 °C for 30 s,
E: 72 °C for 30 s
D: denaturation, A: annealing, E: elongation.
Table A2. Drug physicochemical properties.
Table A2. Drug physicochemical properties.
ParameterValue
Physicochemical properties
Molecular weight (g/mol)431.8
pKa6.27 a
clogD7.45.95 b
clogP6.15 b
a Calculated by a machine learning-based model for pKa prediction [60]. b Predicted by the ACD/LogP program.
Table A3. Summary of monitoring ions and respective parameters for MS detection.
Table A3. Summary of monitoring ions and respective parameters for MS detection.
NameQ1 MassQ3 MassDwell Time (s)Cone (V)Collision (V)
SCR430432.1196.10.154020
Hydroxylated metabolite448.1212.10.155525
Sorafenib (IS)464.9252.20.155538

References

  1. Levy, R.H. Time-dependent pharmacokinetics. Pharmacol. Ther. 1982, 17, 383–397. [Google Scholar] [CrossRef]
  2. Hu, M.H.; Chen, L.J.; Chen, Y.L.; Tsai, M.S.; Shiau, C.W.; Chao, T.I.; Liu, C.Y.; Kao, J.H.; Chen, K.F. Targeting SHP-1-STAT3 signaling: A promising therapeutic approach for the treatment of cholangiocarcinoma. Oncotarget 2017, 8, 65077–65089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Liu, C.Y.; Tseng, L.M.; Su, J.C.; Chang, K.C.; Chu, P.Y.; Tai, W.T.; Shiau, C.W.; Chen, K.F. Novel sorafenib analogues induce apoptosis through SHP-1 dependent STAT3 inactivation in human breast cancer cells. Breast Cancer Res. 2013, 15, R63. [Google Scholar] [CrossRef] [Green Version]
  4. Iwanaga, K.; Honjo, T.; Miyazaki, M.; Kakemi, M. Time-dependent changes in hepatic and intestinal induction of cytochrome P450 3A after administration of dexamethasone to rats. Xenobiotica 2013, 43, 765–773. [Google Scholar] [CrossRef]
  5. Niemi, M.; Backman, J.T.; Fromm, M.F.; Neuvonen, P.J.; Kivisto, K.T. Pharmacokinetic interactions with rifampicin: Clinical relevance. Clin. Pharm. 2003, 42, 819–850. [Google Scholar] [CrossRef] [PubMed]
  6. Baneyx, G.; Parrott, N.; Meille, C.; Iliadis, A.; Lave, T. Physiologically based pharmacokinetic modeling of CYP3A4 induction by rifampicin in human: Influence of time between substrate and inducer administration. Eur. J. Pharm. Sci. 2014, 56, 1–15. [Google Scholar] [CrossRef] [PubMed]
  7. Fahrmayr, C.; Fromm, M.F.; Konig, J. Hepatic OATP and OCT uptake transporters: Their role for drug-drug interactions and pharmacogenetic aspects. Drug Metab. Rev. 2010, 42, 380–401. [Google Scholar] [CrossRef]
  8. Niu, C.R.; Smith, B.; Lai, Y.R. Transporter Gene Regulation in Sandwich Cultured Human Hepatocytes Through the Activation of Constitutive Androstane Receptor (CAR) or Aryl Hydrocarbon Receptor (AhR). Front. Pharmacol. 2021, 11, 620197. [Google Scholar] [CrossRef]
  9. Patel, M.; Taskar, K.S.; Zamek-Gliszczynski, M.J. Importance of Hepatic Transporters in Clinical Disposition of Drugs and Their Metabolites. J. Clin. Pharmacol. 2016, 56, S23–S39. [Google Scholar] [CrossRef] [Green Version]
  10. Kusuhara, H.; Sugiyama, Y. In vitro-in vivo extrapolation of transporter-mediated clearance in the liver and kidney. Drug Metab. Pharm. 2009, 24, 37–52. [Google Scholar] [CrossRef]
  11. Varma, M.V.; Steyn, S.J.; Allerton, C.; El-Kattan, A.F. Predicting Clearance Mechanism in Drug Discovery: Extended Clearance Classification System (ECCS). Pharm. Res. 2015, 32, 3785–3802. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Huo, M.; Zhou, J.; Xie, S. PKSolver: An add-in program for pharmacokinetic and pharmacodynamic data analysis in Microsoft Excel. Comput. Methods Programs Biomed. 2010, 99, 306–314. [Google Scholar] [CrossRef]
  13. Li, R.; Ghosh, A.; Maurer, T.S.; Kimoto, E.; Barton, H.A. Physiologically based pharmacokinetic prediction of telmisartan in human. Drug Metab. Dispos. 2014, 42, 1646–1655. [Google Scholar] [CrossRef] [Green Version]
  14. Yu, S.; Li, S.; Yang, H.; Lee, F.; Wu, J.T.; Qian, M.G. A novel liquid chromatography/tandem mass spectrometry based depletion method for measuring red blood cell partitioning of pharmaceutical compounds in drug discovery. Rapid Commun. Mass Spectrom. 2005, 19, 250–254. [Google Scholar] [CrossRef]
  15. Srivastava, A.; Pike, A.; Williamson, B.; Fenner, K. A Novel Method for Preventing Non-specific Binding in Equilibrium Dialysis Assays Using Solutol® as an Additive. J. Pharm. Sci. 2021, 110, 1412–1417. [Google Scholar] [CrossRef] [PubMed]
  16. Li, N.; Badrinarayanan, A.; Li, X.; Roberts, J.; Hayashi, M.; Virk, M.; Gupta, A. Comparison of In Vitro to In Vivo Extrapolation Approaches for Predicting Transporter-Mediated Hepatic Uptake Clearance Using Suspended Rat Hepatocytes. Drug Metab. Dispos. 2020, 48, 861–872. [Google Scholar] [CrossRef] [PubMed]
  17. Yoshikado, T.; Toshimoto, K.; Nakada, T.; Ikejiri, K.; Kusuhara, H.; Maeda, K.; Sugiyama, Y. Comparison of Methods for Estimating Unbound Intracellular-to-Medium Concentration Ratios in Rat and Human Hepatocytes Using Statins. Drug Metab. Dispos. 2017, 45, 779–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Seglen, P.O. Preparation of isolated rat liver cells. Methods Cell Biol. 1976, 13, 29–83. [Google Scholar] [CrossRef]
  19. Soars, M.G.; Grime, K.; Sproston, J.L.; Webborn, P.J.; Riley, R.J. Use of hepatocytes to assess the contribution of hepatic uptake to clearance in vivo. Drug Metab. Dispos. 2007, 35, 859–865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Musther, H.; Harwood, M.D.; Yang, J.; Turner, D.B.; Rostami-Hodjegan, A.; Jamei, M. The Constraints, Construction, and Verification of a Strain-Specific Physiologically Based Pharmacokinetic Rat Model. J. Pharm. Sci. 2017, 106, 2826–2838. [Google Scholar] [CrossRef] [PubMed]
  21. Watanabe, T.; Kusuhara, H.; Sugiyama, Y. Application of physiologically based pharmacokinetic modeling and clearance concept to drugs showing transporter-mediated distribution and clearance in humans. J. Pharm. Pharm. 2010, 37, 575–590. [Google Scholar] [CrossRef]
  22. Izumi, S.; Nozaki, Y.; Komori, T.; Takenaka, O.; Maeda, K.; Kusuhara, H.; Sugiyama, Y. Comparison of the predictability of human hepatic clearance for organic anion transporting polypeptide substrate drugs between different in vitro–in vivo extrapolation approaches. J. Pharm. Sci. 2017, 106, 2678–2687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Nozaki, Y.; Izumi, S. Recent advances in preclinical in vitro approaches towards quantitative prediction of hepatic clearance and drug-drug interactions involving organic anion transporting polypeptide (OATP) 1B transporters. Drug Metab. Pharm. 2020, 35, 56–70. [Google Scholar] [CrossRef] [PubMed]
  24. Koyanagi, T.; Yano, K.; Kim, S.; Murayama, N.; Yamazaki, H.; Tamai, I. In vivo hepatic clearance of lipophilic drugs predicted by in vitro uptake data into cryopreserved hepatocytes suspended in sera of rats, guinea pigs, monkeys and humans. Xenobiotica 2019, 49, 887–894. [Google Scholar] [CrossRef]
  25. Shitara, Y.; Maeda, K.; Ikejiri, K.; Yoshida, K.; Horie, T.; Sugiyama, Y. Clinical significance of organic anion transporting polypeptides (OATPs) in drug disposition: Their roles in hepatic clearance and intestinal absorption. Biopharm. Drug Dispos. 2013, 34, 45–78. [Google Scholar] [CrossRef]
  26. Li, N.; Badrinarayanan, A.; Ishida, K.; Li, X.; Roberts, J.; Wang, S.; Hayashi, M.; Gupta, A. Albumin-mediated uptake improves human clearance prediction for hepatic uptake transporter substrates aiding a mechanistic in vitro-in vivo extrapolation (IVIVE) strategy in discovery research. AAPS J. 2021, 23, 1. [Google Scholar] [CrossRef] [PubMed]
  27. Roberts, M.S.; Rowland, M. A dispersion model of hepatic elimination: 1. Formulation of the model and bolus considerations. J Pharm. Biopharm. 1986, 14, 227–260. [Google Scholar] [CrossRef]
  28. Poulin, P.; Theil, F.P. Prediction of pharmacokinetics prior to in vivo studies. II. Generic physiologically based pharmacokinetic models of drug disposition. J. Pharm. Sci. 2002, 91, 1358–1370. [Google Scholar] [CrossRef]
  29. Yoshikado, T.; Yoshida, K.; Kotani, N.; Nakada, T.; Asaumi, R.; Toshimoto, K.; Maeda, K.; Kusuhara, H.; Sugiyama, Y. Quantitative Analyses of Hepatic OATP-Mediated Interactions Between Statins and Inhibitors Using PBPK Modeling with a Parameter Optimization Method. Clin Pharm. 2016, 100, 513–523. [Google Scholar] [CrossRef]
  30. Sweeney, L.M.; Gargas, M.L.; Strother, D.E.; Kedderis, G.L. Physiologically based pharmacokinetic model parameter estimation and sensitivity and variability analyses for acrylonitrile disposition in humans. Toxicol. Sci. 2003, 71, 27–40. [Google Scholar] [CrossRef]
  31. Evans, M.V.; Andersen, M.E. Sensitivity analysis of a physiological model for 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD): Assessing the impact of specific model parameters on sequestration in liver and fat in the rat. Toxicol. Sci. 2000, 54, 71–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Mirfazaelian, A.; Kim, K.B.; Anand, S.S.; Kim, H.J.; Tornero-Velez, R.; Bruckner, J.V.; Fisher, J.W. Development of a physiologically based pharmacokinetic model for deltamethrin in the adult male Sprague-Dawley rat. Toxicol. Sci. 2006, 93, 432–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Lin, Z.; Fisher, J.W.; Ross, M.K.; Filipov, N.M. A physiologically based pharmacokinetic model for atrazine and its main metabolites in the adult male C57BL/6 mouse. Toxicol. Appl. Pharm. 2011, 251, 16–31. [Google Scholar] [CrossRef] [PubMed]
  34. Yoon, M.; Nong, A.; Clewell, H.J., 3rd; Taylor, M.D.; Dorman, D.C.; Andersen, M.E. Evaluating placental transfer and tissue concentrations of manganese in the pregnant rat and fetuses after inhalation exposures with a PBPK model. Toxicol. Sci. 2009, 112, 44–58. [Google Scholar] [CrossRef] [Green Version]
  35. Li, M.; Gehring, R.; Riviere, J.E.; Lin, Z. Development and application of a population physiologically based pharmacokinetic model for penicillin G in swine and cattle for food safety assessment. Food Chem. Toxicol. 2017, 107, 74–87. [Google Scholar] [CrossRef]
  36. Gujer, W. Systems Analysis for Water Technology; Springer Science & Business Media: Berlin, Germany, 2008. [Google Scholar]
  37. Oh, J.H.; Lee, J.H.; Lee, Y.J. Evaluation of the Mrp2-mediated flavonoid-drug interaction potential of quercetin in rats and in vitro models. Asian J. Pharm. Sci. 2019, 14, 621–630. [Google Scholar] [CrossRef]
  38. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef]
  39. Kilford, P.J.; Gertz, M.; Houston, J.B.; Galetin, A. Hepatocellular binding of drugs: Correction for unbound fraction in hepatocyte incubations using microsomal binding or drug lipophilicity data. Drug Metab. Dispos. 2008, 36, 1194–1197. [Google Scholar] [CrossRef] [Green Version]
  40. Page, K.M. Validation of Early Human Dose Prediction: A Key Metric for Compound Progression in Drug Discovery. Mol. Pharm. 2016, 13, 609–620. [Google Scholar] [CrossRef]
  41. De Bruyn Kops, C.; Stork, C.; Sicho, M.; Kochev, N.; Svozil, D.; Jeliazkova, N.; Kirchmair, J. GLORY: Generator of the Structures of Likely Cytochrome P450 Metabolites Based on Predicted Sites of Metabolism. Front. Chem. 2019, 7, 402. [Google Scholar] [CrossRef]
  42. Watanabe, R.; Esaki, T.; Kawashima, H.; Natsume-Kitatani, Y.; Nagao, C.; Ohashi, R.; Mizuguchi, K. Predicting Fraction Unbound in Human Plasma from Chemical Structure: Improved Accuracy in the Low Value Ranges. Mol. Pharm. 2018, 15, 5302–5311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Grime, K.; Riley, R.J. The impact of in vitro binding on in vitro-in vivo extrapolations, projections of metabolic clearance and clinical drug-drug interactions. Curr. Drug Metab. 2006, 7, 251–264. [Google Scholar] [CrossRef] [PubMed]
  44. Watanabe, T.; Kusuhara, H.; Maeda, K.; Shitara, Y.; Sugiyama, Y. Physiologically based pharmacokinetic modeling to predict transporter-mediated clearance and distribution of pravastatin in humans. J. Pharm. Exp. 2009, 328, 652–662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Sinz, M.; Wallace, G.; Sahi, J.J.T.A.j. Current industrial practices in assessing CYP450 enzyme induction: Preclinical and clinical. AAPS J. 2008, 10, 391–400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Hakkola, J.; Hukkanen, J.; Turpeinen, M.; Pelkonen, O. Inhibition and induction of CYP enzymes in humans: An update. Arch. Toxicol. 2020, 94, 3671–3722. [Google Scholar] [CrossRef] [PubMed]
  47. Punyawudho, B.; Cloyd, J.C.; Leppik, I.E.; Ramsay, R.E.; Marino, S.E.; Pennell, P.B.; White, J.R.; Birnbaum, A.K. Characterization of the time course of carbamazepine deinduction by an enzyme turnover model. Clin. Pharmacokinet. 2009, 48, 313–320. [Google Scholar] [CrossRef] [PubMed]
  48. Perdomo, V.G.; Rigalli, J.P.; Villanueva, S.S.; Ruiz, M.L.; Luquita, M.G.; Echenique, C.G.; Catania, V.A. Modulation of biotransformation systems and ABC transporters by benznidazole in rats. Antimicrob. Agents Chemother. 2013, 57, 4894–4902. [Google Scholar] [CrossRef] [Green Version]
  49. Dixit, V.; Hariparsad, N.; Li, F.; Desai, P.; Thummel, K.E.; Unadkat, J.D. Cytochrome P450 enzymes and transporters induced by anti-human immunodeficiency virus protease inhibitors in human hepatocytes: Implications for predicting clinical drug interactions. Drug Metab. Dispos. 2007, 35, 1853–1859. [Google Scholar] [CrossRef] [Green Version]
  50. El-Kattan, A.F.; Varma, M.V.S. Navigating Transporter Sciences in Pharmacokinetics Characterization Using the Extended Clearance Classification System. Drug Metab. Dispos. 2018, 46, 729–739. [Google Scholar] [CrossRef]
  51. Wongwan, T.; Chatsudthipong, V.; Soodvilai, S. Farnesoid X Receptor Activation Stimulates Organic Cations Transport in Human Renal Proximal Tubular Cells. Int. J. Mol. Sci. 2020, 21, 6078. [Google Scholar] [CrossRef]
  52. Hyrsova, L.; Smutny, T.; Trejtnar, F.; Pavek, P. Expression of organic cation transporter 1 (OCT1): Unique patterns of indirect regulation by nuclear receptors and hepatospecific gene regulation. Drug Metab. Rev. 2016, 48, 139–158. [Google Scholar] [CrossRef] [PubMed]
  53. Palaiokostas, M.; Ding, W.; Shahane, G.; Orsi, M. Effects of lipid composition on membrane permeation. Soft Matter 2018, 14, 8496–8508. [Google Scholar] [CrossRef] [Green Version]
  54. Riccardi, K.; Ryu, S.; Lin, J.; Yates, P.; Tess, D.; Li, R.; Singh, D.; Holder, B.R.; Kapinos, B.; Chang, G.; et al. Comparison of Species and Cell-Type Differences in Fraction Unbound of Liver Tissues, Hepatocytes, and Cell Lines. Drug Metab. Dispos. 2018, 46, 415–421. [Google Scholar] [CrossRef] [PubMed]
  55. Xu, E.; Forest, M.P.; Schwab, M.; Avramoglu, R.K.; St-Amand, E.; Caron, A.Z.; Bellmann, K.; Shum, M.; Voisin, G.; Paquet, M.; et al. Hepatocyte-specific Ptpn6 deletion promotes hepatic lipid accretion, but reduces NAFLD in diet-induced obesity: Potential role of PPARgamma. Hepatology 2014, 59, 1803–1815. [Google Scholar] [CrossRef] [PubMed]
  56. Hirano, M.; Maeda, K.; Shitara, Y.; Sugiyama, Y. Contribution of OATP2 (OATP1B1) and OATP8 (OATP1B3) to the hepatic uptake of pitavastatin in humans. J. Pharm. Exp. 2004, 311, 139–146. [Google Scholar] [CrossRef]
  57. Giacomini, K.M.; Huang, S.M.; Tweedie, D.J.; Benet, L.Z.; Brouwer, K.L.R.; Chu, X.Y.; Dahlin, A.; Evers, R.; Fischer, V.; Hillgren, K.M.; et al. Membrane transporters in drug development. Nat. Rev. Drug Discov. 2010, 9, 215–236. [Google Scholar] [CrossRef]
  58. Jones, H.M.; Barton, H.A.; Lai, Y.; Bi, Y.-a.; Kimoto, E.; Kempshall, S.; Tate, S.C.; El-Kattan, A.; Houston, J.B.; Galetin, A.; et al. Mechanistic pharmacokinetic modeling for the prediction of transporter-mediated disposition in humans from sandwich culture human hepatocyte data. Drug Metab. Dispos. 2012, 40, 1007–1017. [Google Scholar] [CrossRef] [Green Version]
  59. Watanabe, T.; Kusuhara, H.; Maeda, K.; Kanamaru, H.; Saito, Y.; Hu, Z.; Sugiyama, Y. Investigation of the rate-determining process in the hepatic elimination of HMG-CoA reductase inhibitors in rats and humans. Drug Metab. Dispos. 2010, 38, 215–222. [Google Scholar] [CrossRef] [Green Version]
  60. Baltruschat, M.; Czodrowski, P. Machine learning meets pKa. F1000Res 2020, 9, 113. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the experimental procedure’s schedule. In this study, rats were divided into three groups. (1) Control group (red-colored, n = 6): the rats were not given a vehicle or drug; (2) Vehicle group (blue-colored, n = 6): the rats were given a vehicle once a day orally for 2 weeks; (3) Repeated-dosing group (black-colored, n = 6): the rats were given 30 mg/kg of the drug dissolved in the vehicle once a day orally for 2 weeks. After a washout period of two days, the SCR430 that was dissolved in the vehicle was administered intravenously or intraperitoneally at a dose of 3 mg/kg for each group.
Figure 1. Schematic representation of the experimental procedure’s schedule. In this study, rats were divided into three groups. (1) Control group (red-colored, n = 6): the rats were not given a vehicle or drug; (2) Vehicle group (blue-colored, n = 6): the rats were given a vehicle once a day orally for 2 weeks; (3) Repeated-dosing group (black-colored, n = 6): the rats were given 30 mg/kg of the drug dissolved in the vehicle once a day orally for 2 weeks. After a washout period of two days, the SCR430 that was dissolved in the vehicle was administered intravenously or intraperitoneally at a dose of 3 mg/kg for each group.
Pharmaceutics 14 02562 g001
Figure 2. (a,b) Mean plasma concentration–time curve of SCR430 after IV (a) or IP (b) administration at a dose of 3 mg kg−1 (△, control (without a dosing vehicle or SCR430) group; □, vehicle (vehicle only, 14 days, PO) group; ●, repeated-dosing (SCR430 30 mg/kg, 14 days, PO) group (n = 6). (ce) Comparison of the control, vehicle, and repeated-dosing groups’ pharmacokinetic parameters. The AUC from time 0 h to infinity (AUC0–∞) (c), the total plasma clearance (CL) (d), and the steady-state volume of distribution (Vdss) (e) were compared. A one-way ANOVA statistical analysis was used for the statistical analysis, followed by a Turkey’s post-hoc test, where appropriate. The stars indicate statistical significances, ** p < 0.01, *** p < 0.001.
Figure 2. (a,b) Mean plasma concentration–time curve of SCR430 after IV (a) or IP (b) administration at a dose of 3 mg kg−1 (△, control (without a dosing vehicle or SCR430) group; □, vehicle (vehicle only, 14 days, PO) group; ●, repeated-dosing (SCR430 30 mg/kg, 14 days, PO) group (n = 6). (ce) Comparison of the control, vehicle, and repeated-dosing groups’ pharmacokinetic parameters. The AUC from time 0 h to infinity (AUC0–∞) (c), the total plasma clearance (CL) (d), and the steady-state volume of distribution (Vdss) (e) were compared. A one-way ANOVA statistical analysis was used for the statistical analysis, followed by a Turkey’s post-hoc test, where appropriate. The stars indicate statistical significances, ** p < 0.01, *** p < 0.001.
Pharmaceutics 14 02562 g002
Figure 3. (a) Comparison of the total (light gray, n = 6), biliary (dark gray, n = 3), and urinary (black, n = 3) SCR430 clearance after IV SCR430 bolus administration (3 mpk) in the control, vehicle, and repeated-dosing groups. The biliary clearance (CLb) (b) and urinary clearance (CLr) (c) were also compared. A one-way ANOVA statistical analysis was used for the statistical analysis, followed by a Turkey’s post-hoc test, where appropriate. The stars indicate statistical significances, ** p < 0.01.
Figure 3. (a) Comparison of the total (light gray, n = 6), biliary (dark gray, n = 3), and urinary (black, n = 3) SCR430 clearance after IV SCR430 bolus administration (3 mpk) in the control, vehicle, and repeated-dosing groups. The biliary clearance (CLb) (b) and urinary clearance (CLr) (c) were also compared. A one-way ANOVA statistical analysis was used for the statistical analysis, followed by a Turkey’s post-hoc test, where appropriate. The stars indicate statistical significances, ** p < 0.01.
Pharmaceutics 14 02562 g003
Figure 4. (a) Three candidate metabolites of the SCR430 aromatic hydroxyl site via CYP-mediated metabolism predicted by metabolite profiling. (b) Representative UPLC-qToF-MS chromatograms for the two hydroxylated SCR430 metabolite types in the liver tissue, plasma, bile, and hepatocytes of the repeated-dosing group. (c) The equivalent cumulative percentage–time profile of the dose recovered as an aromatic hydroxylated metabolite in the bile after IV administration among the control, vehicle, and repeated-dosing groups. (Data = mean ± standard deviation (SD), n = 3).
Figure 4. (a) Three candidate metabolites of the SCR430 aromatic hydroxyl site via CYP-mediated metabolism predicted by metabolite profiling. (b) Representative UPLC-qToF-MS chromatograms for the two hydroxylated SCR430 metabolite types in the liver tissue, plasma, bile, and hepatocytes of the repeated-dosing group. (c) The equivalent cumulative percentage–time profile of the dose recovered as an aromatic hydroxylated metabolite in the bile after IV administration among the control, vehicle, and repeated-dosing groups. (Data = mean ± standard deviation (SD), n = 3).
Pharmaceutics 14 02562 g004
Figure 5. (ac) The degradation rate of SCR430 in the suspended hepatocytes from the three groups of rats: control (a), vehicle (b), and repeated-dosing (c) groups. (d,e) Integration plots for the hepatic uptake clearance of SCR430 in the control (d) and repeated-dosing (e) groups at the condition of 4 °C (open) and 37 °C (close). The amount of SCR430 within the hepatocytes divided by the concentration (t) in the buffer was plotted against the AUC0–∞ (buff) divided by the concentration (t) in the buffer. The solid lines represent the linear regression line for the control and repeated-dosing groups. (f,g) The Kp (C/M ratio) of SCR430 was assessed at 4 °C (open) and 37 °C (close) to determine the Kpuu,ss in the control (f) and repeated-dosing groups (g) (Data = mean ± SD).
Figure 5. (ac) The degradation rate of SCR430 in the suspended hepatocytes from the three groups of rats: control (a), vehicle (b), and repeated-dosing (c) groups. (d,e) Integration plots for the hepatic uptake clearance of SCR430 in the control (d) and repeated-dosing (e) groups at the condition of 4 °C (open) and 37 °C (close). The amount of SCR430 within the hepatocytes divided by the concentration (t) in the buffer was plotted against the AUC0–∞ (buff) divided by the concentration (t) in the buffer. The solid lines represent the linear regression line for the control and repeated-dosing groups. (f,g) The Kp (C/M ratio) of SCR430 was assessed at 4 °C (open) and 37 °C (close) to determine the Kpuu,ss in the control (f) and repeated-dosing groups (g) (Data = mean ± SD).
Pharmaceutics 14 02562 g005
Figure 6. In vitro predicted hepatic clearance correlation with the observed hepatic clearance in the control (open triangle) and repeated-dosing (closed circle) groups. (a) Method 1: Predicted hepatic clearance based on the intrinsic metabolic clearance (CLint,met). (b) Method 2: Predicted hepatic clearance based on the hepatic uptake clearance (PSinf). (c) Method 3: Predicted hepatic clearance based on the corrected CLint,met by multiplying with Kpuu,ss. The solid lines represent the line of unity, which is the best fit, and the dashed lines represent the 3-fold error from the unity.
Figure 6. In vitro predicted hepatic clearance correlation with the observed hepatic clearance in the control (open triangle) and repeated-dosing (closed circle) groups. (a) Method 1: Predicted hepatic clearance based on the intrinsic metabolic clearance (CLint,met). (b) Method 2: Predicted hepatic clearance based on the hepatic uptake clearance (PSinf). (c) Method 3: Predicted hepatic clearance based on the corrected CLint,met by multiplying with Kpuu,ss. The solid lines represent the line of unity, which is the best fit, and the dashed lines represent the 3-fold error from the unity.
Pharmaceutics 14 02562 g006
Figure 7. Schematic diagram of the mPBPK model with a single-adjustment compartment was designed to predict the plasma and liver concentrations in the control and repeated-dosing groups. The liver unit was composed of five units to mimic the dispersion model. The dotted, dashed line between the inlet and liver units is the passive diffusion clearance (PSdiff), and the solid downward line is the active transporter-mediated uptake (PSinf,act) clearance. The solid downward line at the bottom of the liver units is the metabolic clearance (CLint,met).
Figure 7. Schematic diagram of the mPBPK model with a single-adjustment compartment was designed to predict the plasma and liver concentrations in the control and repeated-dosing groups. The liver unit was composed of five units to mimic the dispersion model. The dotted, dashed line between the inlet and liver units is the passive diffusion clearance (PSdiff), and the solid downward line is the active transporter-mediated uptake (PSinf,act) clearance. The solid downward line at the bottom of the liver units is the metabolic clearance (CLint,met).
Pharmaceutics 14 02562 g007
Figure 8. Simulated and observed intravenous plasma concentration–time profiles of SCR430 in the control and repeated-dosing groups. The Monte Carlo simulation curves (shaded region) and the observed data for the SCR430 concentrations in the plasma and liver of the control (open triangle; (a,c)) and repeated-dosing (closed circle; (b,d)) groups after intravenous administration at a dose of 3 mg kg−1. The shaded region represents 95% confidence intervals around the 5th and 95th (light gray-shaded region), 50th (green or yellow dotted line), 20th and 80th (middle gray-shaded region), and 35th and 65th (deep gray-shaded region) percentiles of the simulated concentration. The observed plasma concentrations are represented by means (symbols) and standard deviations (bars). The observed liver concentration is represented by a mean value.
Figure 8. Simulated and observed intravenous plasma concentration–time profiles of SCR430 in the control and repeated-dosing groups. The Monte Carlo simulation curves (shaded region) and the observed data for the SCR430 concentrations in the plasma and liver of the control (open triangle; (a,c)) and repeated-dosing (closed circle; (b,d)) groups after intravenous administration at a dose of 3 mg kg−1. The shaded region represents 95% confidence intervals around the 5th and 95th (light gray-shaded region), 50th (green or yellow dotted line), 20th and 80th (middle gray-shaded region), and 35th and 65th (deep gray-shaded region) percentiles of the simulated concentration. The observed plasma concentrations are represented by means (symbols) and standard deviations (bars). The observed liver concentration is represented by a mean value.
Pharmaceutics 14 02562 g008
Figure 9. Local sensitivity analysis for the mPBPK modeling in the liver and plasma of the control group.
Figure 9. Local sensitivity analysis for the mPBPK modeling in the liver and plasma of the control group.
Pharmaceutics 14 02562 g009
Figure 10. The relative expression of representative nuclear receptors, hepatic uptake transporters, and metabolic enzymes in the liver. The repeated-dosing group mRNA levels were expressed as a relative fold change compared to the control group (means ± the standard error of the mean). A Student’s t-test was employed to detect significant differences in the mRNA expression levels in the tissues (** p < 0.01).
Figure 10. The relative expression of representative nuclear receptors, hepatic uptake transporters, and metabolic enzymes in the liver. The repeated-dosing group mRNA levels were expressed as a relative fold change compared to the control group (means ± the standard error of the mean). A Student’s t-test was employed to detect significant differences in the mRNA expression levels in the tissues (** p < 0.01).
Pharmaceutics 14 02562 g010
Table 1. Comparison of the pharmacokinetic parameters between the control (without a dosing vehicle or SCR430, n = 6), vehicle (vehicle only, 14 days, PO, n = 6), and repeated-dosing (SCR430 30 mpk, 14 days, PO, n = 6) groups. The data are presented as mean ± standard deviation (relative standard error %).
Table 1. Comparison of the pharmacokinetic parameters between the control (without a dosing vehicle or SCR430, n = 6), vehicle (vehicle only, 14 days, PO, n = 6), and repeated-dosing (SCR430 30 mpk, 14 days, PO, n = 6) groups. The data are presented as mean ± standard deviation (relative standard error %).
ParameterControl GroupVehicle GroupRepeated-Dosing Group
IV (3 mg/kg)
AUC0→∞ (μg/mL·min)6333 ± 1505 (9.7%)6319 ± 2854 (18.4%)1420 ± 446 ** (12.8%)
CL (mL/h/kg)30.02 ± 8.29 (11.3%) 32.99 ± 12.54 (15.4%)137.9 ± 43.8 *** (12.9%)
Vdss (mL/kg)286.4 ± 65 (9.2%)316.0 ± 58.7 (7.6%)378.4 ± 119.3 (12.8%)
CLb (mL/h/kg)0.05 ± 0.02 (22%)0.061 ± 0.01 (9.8%)0.075 ± 0.003 (1.3%)
CLr (mL/h/kg)0.005 ± 0.002 (20%)0.003 ± 0.002 (33%)0.007 ± 0.006 (42%)
Rb3.95 ± 0.64 (9.4%)4.58 ± 0.47 (5.9%)4.53 ± 0.65 (8.2%)
IP (3 mg/kg)
AUCIP,0→10h (ng/mL·h)7604 ± 3110 (16.7%)6894 ± 1004 (5.9%)951 ± 302 *** (12.9%)
Binding properties
Fraction unbound (fu)
fu,p a0.0078 ± 0.001 (7.7%)0.0074 ± 0.002 (13.5%)0.0081 ± 0.002 (12.3%)
fu,inc,hepa b0.032 ± 0.004 (6.2%) Equal to control
Predicted fu,inc,heap c0.024 0.024
Extrapolated
fu,inc,hepa d

0.088

0.088
fu,hepa  e0.004 ± 0.002 (25%) 0.01 ± 0.008 (40%)
Tissue:Plasma
Partition Coefficients (Kp)
Kp_liver2.06 ± 0.06 (1.4%)1.27 ± 0.15 (6.3%)1.96 ± 1.09 (32.1%)
Kp_kidney1.20 ± 0.1 (5%)1.37 ± 0.15 (5.8%)5.50 ± 0.82 * (8.5%)
a determined experimentally. b determined experimentally. c predicted by an in silico method [39]. d extrapolated from plasma protein binding [40]. e determined by a Kp method at 4 °C in 120 min. * p < 0.05, ** p < 0.01, *** p < 0.001.
Table 2. Predicted and observed hepatic clearance calculated using IVIVE Methods 1, 2, and 3.
Table 2. Predicted and observed hepatic clearance calculated using IVIVE Methods 1, 2, and 3.
UnitControlVehicleRepeated-Dosing
Method 1
in vitro intrinsic metabolic clearance (CLint,met,vitro)μL/min per 1061.321.211.69
in vivo overall intrinsic clearance (CLint,all)mL/min per kg160.4147.0205.3
Predicted hepatic clearance (CLh)mL/h per kg74.868.695.7
Method 2
in vitro hepatocyte uptake clearance (PSinf,hep)μL/min per 10611.9 43.2
in vivo overall intrinsic clearance (CLint,all)mL/min per kg46.2 168.0
Predicted hepatic clearance (CLh)mL/h per kg21.6 78.4
Method 3
intrinsic metabolic clearance (CLint,met)mL/min per kg160.4 205.3
liver-to-plasma unbound drug concentration ratio at steady state (Kpuu,ss) 0.11 1.7
in vivo overall intrinsic clearance (CLint,all)mL/min per kg17.3 344.3
Predicted hepatic clearance (CLh)mL/h per kg8.1 160
Observed in vivo clearance (CL)mL/h per kg30.02 32.99137.9
Table 3. Input parameter values used for the mPBPK modeling and Monte Carlo simulation.
Table 3. Input parameter values used for the mPBPK modeling and Monte Carlo simulation.
Parameters Comment
ControlRepeated-Dosing
Single-adjustment compartment
kin (hr−1)2.7Equal to ControlFitted from observed data
kout (hr−1)1.0Equal to ControlFitted from observed data
Tissue volume (L/kg)
Central 0.018Equal to ControlFitted from observed data
Liver 0.037Equal to Control[28]
Extracellular space in the liver0.01Equal to Control[44]
Hepatocytes0.027Equal to Control[44]
Blood flow (L/h/kg)
Liver 3.69Equal to Control[28]
Monte Carlo simulation parameters
PSinf,act (L/h/kg)(2.5, 0.8)(9.2, 0.6)(Mean, SD)
PSdiff (L/h/kg)(0.3, 0.09)(0.86, 0.4)(Mean, SD)
CLint,met (L/h/kg)(9.8, 1.0)(12.3, 0.08)(Mean, SD)
fu,hepa(0.004, 0.002)(0.01, 0.008)(Mean, SD)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, M.-C.; Lee, Y.-J. Analysis of Time-Dependent Pharmacokinetics Using In Vitro–In Vivo Extrapolation and Physiologically Based Pharmacokinetic Modeling. Pharmaceutics 2022, 14, 2562. https://doi.org/10.3390/pharmaceutics14122562

AMA Style

Kim M-C, Lee Y-J. Analysis of Time-Dependent Pharmacokinetics Using In Vitro–In Vivo Extrapolation and Physiologically Based Pharmacokinetic Modeling. Pharmaceutics. 2022; 14(12):2562. https://doi.org/10.3390/pharmaceutics14122562

Chicago/Turabian Style

Kim, Min-Chang, and Young-Joo Lee. 2022. "Analysis of Time-Dependent Pharmacokinetics Using In Vitro–In Vivo Extrapolation and Physiologically Based Pharmacokinetic Modeling" Pharmaceutics 14, no. 12: 2562. https://doi.org/10.3390/pharmaceutics14122562

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop